Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

A regioregular terpolymer comprising two electron-deficient and one electron-rich unit for ultra small band gap solar cells

Weiwei Li *ab, Koen H. Hendriks bc, Alice Furlan bc, Andong Zhang a, Martijn M. Wienk bc and René A. J. Janssen *bc
aBeijing National Laboratory for Molecular Sciences, CAS Key Laboratory of Organic Solids, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, P. R. China. E-mail: liweiwei@iccas.ac.cn
bMolecular Materials and Nanosystems, Institute of Complex Molecular Systems, Eindhoven University of Technology, P. O. Box 513, 5600 MB Eindhoven, The Netherlands. E-mail: r.a.j.janssen@tue.nl
cDutch Institute for Fundamental Energy Research, De Zaale 20, 5612 AJ Eindhoven, The Netherlands

Received 27th December 2014 , Accepted 3rd February 2015

First published on 5th February 2015


Abstract

A regioregular terpolymer that combines electron-deficient diketopyrrolopyrrole and benzothiadiazole units with an electron-rich dithienopyrrole unit is presented. In solar cells, the terpolymer affords an open circuit voltage of 0.52 V and a power conversion efficiency of 3.7% with a spectral response up to 1050 nm.


Bulk heterojunction polymer solar cells (PSCs) with conjugated polymers as electron donor and fullerene derivatives such as [6,6]-phenyl-C71-butyric acid methyl ester ([70]PCBM) as electron acceptor have significantly progressed in the last decade and now reach power conversion efficiencies (PCEs) above 10% in single junction solar cells.1–3 The efficiencies have been further increased to 11–12% via tandem and triple junction device architectures,4,5 in which two or three photoactive layers with complementary absorption spectra are stacked and connected in series. By applying different conjugated polymers that absorb light from the visible light to the near-infrared region above 1000 nm, the conversion efficiencies of triple junction solar cells may achieve 15–20%.6 To reach this ambitious goal, it will be important to develop efficient conjugated polymers that absorb up to 1000 nm and beyond.

Conjugated polymers with tailored optical band gaps (Eg) can be obtained using electron donating and electron deficient units alternating along the main chain. In general stronger electron donors and stronger acceptors reduce the optical band gap. Using this design motif and by employing suitable side chains, the energy levels, solubility, aggregation and crystallinity of the polymers can be tuned. Alternating conjugated polymers have already reached high performance in polymer solar cells, especially for wide (Eg > 1.7 eV) and medium (1.7 eV > Eg > 1.4 eV) optical band gaps with PCEs above 9%.1–3 A few small band gap conjugated polymers with Eg close to 1.4 eV provide ≥8% PCE.7–9 For Eg below 1.25 eV the most efficient cells reach 5.3% PCE.10 In many of the successful materials a significant photon energy loss between Eg and open circuit voltage (Voc) is incurred, Eloss = EgqVoc. Eloss provides a driving force for photoinduced charge generation and often enhances the short circuit current density (Jsc). When the optical band gap is reduced, charge generation is usually negligible when Eloss becomes less than 0.6 eV.11 This problem becomes especially important when the absorption of conjugated polymers extends above 1000 nm where the optical band gap becomes less than 1.2 eV. To maintain a high quantum efficiency for charge generation, the Voc, which is controlled by the energy difference between the HOMO of the donor and the LUMO of the acceptor, must be designed to be less than 0.6 V. However, a too low Voc would quickly lead to an insignificant PCE. Hence, conjugated polymers with a very small band gap require a judicious design in terms of energy levels to be applicable in PSCs.

Several conjugated polymers with absorption beyond 1000 nm have been reported, often using strong electron deficient units such as thienoisoindigo,12,13 thiadiazoloquinoxaline,14,15 bisbenzo thiadiazole,16 pyrazinoquinoxaline17,18 or diketopyrrolopyrrole (DPP).19 When applied in polymer solar cells PCEs up to 5.3% have been obtained.10 Especially, PDPP2T-DTP (Fig. 1a) with DPP as electron acceptor and dithienopyrrole (DTP) as strong electron donor affords Eg = 1.23 eV19 and provides Jsc > 20 mA cm−2.10 The Voc of cells based on this polymer is, however, relatively low with 0.43 V, resulting in Eloss = 0.80 eV. In order to reduce Eloss and enhance Voc, we introduce the use of two electron deficient units, DPP and benzothiadiazole (BT), in combination with DTP as electron donor. The new ultra low band gap polymer, PDPP2T-BT-co-DTP (Fig. 1c) provides a significantly higher Voc of 0.52 eV and PCE of 3.7% with a reduced optical band gap of 1.19 eV in polymer solar cells.


image file: c4cc10357b-f1.tif
Fig. 1 The DPP polymers with (a) dithienopyrrole (DTP), (b) benzothiadiazole (BT) and (c) both DTP and BT.

PDPP2T-DTP10 and PDPP2T-BT20 were synthesized following previously reported procedures. The regioregular alternating copolymer PDPP2T-BT-co-DTP was synthesized using the dibromo-monomer of T-DPP-T-BT-T-DPP-T and the bisstannyl-DTP monomer via Stille polymerization (Scheme S1, ESI). The conditions applied in the polymerization reaction, toluene/DMF 10[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) as solvent and Pd2(dba)3/PPh3 1[thin space (1/6-em)]:[thin space (1/6-em)]4 (mol/mol) as catalyst, were chosen to reduce homo coupling defects and achieve a high molecular weight.21 The viscous gel formed during polymerization indicates a high molecular weight and the work-up required the use of 1,1,2,2-tetrachloroethane at 140 °C. It was not possible to measure the molecular weight by gel permeation chromatography (GPC) in o-dichlorobenzene (o-DCB) because the polymer seems to adhere to or decompose on the column. The same problem was previously encountered for PDPP2T-DTP.10 The thermal stability of PDPP2T-BT-co-DTP is somewhat less than that of PDPP2T-BT and PDPP2T-DTP (Fig. S1, ESI). The strong electron donating DTP unit makes polymers based on DTP susceptible to oxidation in air, and processing of devices was performed under inter atmosphere.10

All three polymers absorb in the near-infrared region with onsets above 1000 nm (Fig. 2a). The absorption in thin films is slightly red-shifted compared to chloroform solution (Fig. S2, ESI). PDPP2T-BT has a lower band gap (1.19 eV) than PDPP2T-DTP (Eg = 1.23 eV). While we anticipated that the combination of two acceptors (DPP and BT) would increase the electron withdrawing character and could further reduce the band gap, the terpolymer PDPP2T-BT-co-DTP, that combines DPP, BT, and DTP also has Eg = 1.19 eV.


image file: c4cc10357b-f2.tif
Fig. 2 (a) Optical absorption spectra of the polymers in solid state films and (b) cyclic voltammograms of the polymers thin films on ITO substrates.

Importantly, however, the introduction of two acceptors allows to effectively tune the frontier orbital energy levels.22,23 Cyclic voltammetry (Fig. 2b and Table 1) reveals that as a consequence of incorporating two electron deficient units (DPP and BT), the PDPP2T-BT copolymer has deep HOMO and LUMO levels (−5.55 eV and −4.01 eV). Hence, the LUMO–LUMO offset (ΔELL) between PDPP2T-BT and [70]PCBM is only 0.15 eV and significantly less than the ∼0.3 eV threshold for charge separation. This problem often occurs in very small band gap conjugated polymers.24,25 With one electron deficient (DPP) and one electron rich unit (DTP), the PDPP2T-DTP copolymer exhibits significantly less negative HOMO and LUMO levels (−5.26 eV and −3.68 eV) to give ΔELL = 0.48 eV. The new terpolymer PDPP2T-BT-co-DTP has the HOMO and LUMO in between those of the two copolymers PDPP2T-BT and PDPP2T-DTP, so that ΔELL = 0.24 eV. The low lying HOMO level will help to achieve a higher Voc than that of PDPP2T-DTP, while the higher ΔELL than that of PDPP2T-BT is beneficial for the efficiency of photoinduced charge separation.

Table 1 Optical and electrochemical properties of the polymers
Polymer

image file: c4cc10357b-t1.tif

(eV)
E filmg (eV) E LUMO (eV) E HOMO (eV) E cvg (eV) ΔELLb (eV)
a Determined using a work function value of −5.23 eV for Fc/Fc+. b ΔELL = q(ELUMO([70]PCBM) − ELUMO) with ELUMO([70]PCBM) = −4.16 V.
PDPP2T-BT 1.22 1.19 −4.01 −5.55 1.54 0.15
PDPP2T-DTP 1.26 1.23 −3.68 −5.26 1.58 0.48
PDPP2T-BT-co-DTP 1.21 1.19 −3.92 −5.35 1.43 0.24


The three polymers were applied in polymer solar cells with a transparent ITO/poly(3,4-ethylenedioxy thiophene):poly(styrene sulfonate) (PEDOT:PSS) bottom contact for hole collection and LiF/Al top contact for electron collection. The photoactive layers, in which the polymers were blended with [70]PCBM, were carefully optimized for the donor to acceptor weight ratio, solvent, co-solvent, and thickness. In general, the optimized donor–acceptor ratio is 1[thin space (1/6-em)]:[thin space (1/6-em)]2 (w/w) and the optimal thickness was around 110 nm, when spin coating from chloroform solution. For PDPP2T-BT 10% o-DCB as co-solvent gave the best performance. For PDPP2T-DTP and PDPP2T-BT-co-DTP the best co-solvents concentrations are 5% o-DCB and 3% 1-chloronaphthalene (1-CN), respectively. The JV characteristics and external quantum efficiencies (EQEs) of the optimized devices are depicted in Fig. 3 and the device parameters are summarized in Table 2.


image file: c4cc10357b-f3.tif
Fig. 3 (a) JV characteristics in dark (dashed lines) and under white light illumination (solid lines) of optimized solar cells of the polymers with [70]PCBM. (b) EQE of the same devices.
Table 2 Solar cell parameters of optimized solar cells of the DPP polymers with [70]PCBM
Polymer J sc (mA cm−2) V oc (V) FF PCEa (%) E loss (eV)
a J sc and PCE were calculated by integrating the EQE spectrum with the AM1.5G spectrum.
PDPP2T-BT 2.8 0.66 0.66 1.2 0.53
PDPP2T-DTP 16.6 0.43 0.54 3.9 0.80
PDPP2T-BT-co-DTP 12.2 0.52 0.58 3.7 0.67


As expected from the low ΔELL the PDPP2T-BT copolymer gives a low PCE of 1.2%, mainly due to a low short circuit current density (Jsc = 2.8 mA cm−2). The cells have a high Voc of 0.66 V at this low band gap (1.19 eV). The photon energy loss Eloss equals 0.53 eV and is below the threshold of 0.6 eV for efficient charge transfer,11 rationalizing the low Jsc. In contrast, photovoltaic devices based on the copolymer PDPP2T-DTP give a high Jsc of 16.6 mA cm−2 and PCE of 3.9%, but a relatively low Voc of 0.43 V as consequence of the higher lying HOMO. The low Voc increases the photon energy loss to Eloss = 0.80 eV. The new terpolymer PDPP2T-BT-co-DTP gives Voc = 0.52 V, i.e. in between the Voc's obtained with the two copolymers, but a somewhat low Jsc (12.2 mA cm−2) and PCE of 3.7%. Here Eloss is 0.67 eV. The differences in photocurrent are also reflected in the EQE spectra (Fig. 3b). In PDPP2T-BT:[70]PCBM cells the contribution of light absorbed by the polymer in the 700–1050 nm region is negligible and photocurrent is only created via excitation of [70]PCBM. In contrast, a broad spectral response from 350 to 1050 nm, covering the absorption spectra of fullerene and polymer is found for PDPP2T-BT-co-DTP:[70]PCBM blends. Unfortunately, the PDPP2T-BT-co-DTP terpolymer provides a relatively low EQE (maximizing at 0.32) compared to that of the PDPP2T-DTP copolymer (maximizing at 0.53).

As previously reported, a lower energy loss decreases the driving force for electron transfer and correlates with a lower EQE.7 Hence, the lower EQE of the terpolymer PDPP2T-BT-co-DTP is possibly related to the reduced ΔELL and Eloss. However, the energy loss of the terpolymer is 0.67 eV, i.e. above the threshold of 0.6 eV.11 Actually, another DPP-polymer, PDPP3T, where Eloss = 0.65 eV gave a maximum EQE of 0.50,7 which is significantly higher than the 0.32 of the terpolymer. Therefore other effects than Eloss likely contribute to the low EQE.

First, TEM reveals that the morphology of the PDPP2T-BT-co-DTP:[70]PCBM blends consists of semicrystalline polymer fibrils (Fig. 4). The crystallinty can be readily seen from the lattice fringes in Fig. 4b. The width of these semicrystalline fibrils exceeds 10 nm, commonly considered as the maximum exciton diffusion length. We have shown recently for a range of DPP-polymers that the fibril width anti-correlates with the EQE and that exciton loss via intrinsic decay may contribute to a low EQE.26


image file: c4cc10357b-f4.tif
Fig. 4 Bright field TEM images of the optimized PDPP2T-BTD-co-DTP:[70]PCBM blend films. Scale bars: (a) 200 nm and (b) 50 nm.

Second, localization of excitons on the T-DPP-T-BT-T-DPP-T segments may occur. Because PDPP2T-BT has a smaller band gap than PDPP2T-DTP, excitons may prefer T-DPP-T-BT-T-DPP-T segments. Excitons on such segments may have a lower tendency to transfer an electron to the fullerene, as suggested by the fact that the PDPP2T-BT copolymer is unable to create charges efficiently. This explanation is supported by our recent work on the detrimental effect of low-LUMO chain segments on the charge generation efficiency.21

In conclusion, a new conjugated terpolymer with an ultra small band gap, Eg = 1.19 eV, was designed and applied in polymer solar cells. By incorporating two electron-deficient and one electron-rich unit into the conjugated backbone the frontier energy levels were tailored to maximize Voc and Jsc. The terpolymer provided a relatively high Voc (0.52 V) and PCE (3.7%) with a photoresponse up to 1050 nm, despite a moderate EQE. The study shows that by extending the push–pull approach and incorporating more units than single electron rich and electron poor units is viable way to create semiconducting polymers with finely tuned energy levels.

The work was performed in the framework of the Large cells and X10D projects that received funding from the European Commission's Seventh Framework Programme (Grant Agreement No. 261936 and No. 287818). The research forms part of the Solliance OPV programme and has received funding from the Ministry of Education, Culture and Science (Gravity program 024.001.035).

Notes and references

  1. Y. Liu, J. Zhao, Z. Li, C. Mu, W. Ma, H. Hu, K. Jiang, H. Lin, H. Ade and H. Yan, Nat. Commun., 2014, 5, 5293 CrossRef CAS PubMed.
  2. J.-D. Chen, C. Cui, Y.-Q. Li, L. Zhou, Q.-D. Ou, C. Li, Y. Li. and J.-X. Tang, Adv. Mater., 2015, 27, 1035–1041 CrossRef PubMed.
  3. S.-H. Lia, H.-J. Jhuo, P.-N. Yeh, Y.-S. Cheng, Y.-L. Li, Y.-H. Lee, S. Sharma and S.-A. Chen, Sci. Rep., 2014, 4, 6813 CrossRef PubMed.
  4. C.-C. Chen, W.-H. Chang, K. Yoshimura, K. Ohya, J. You, J. Gao, Z. Hong and Y. Yang, Adv. Mater., 2014, 26, 5670–5677 CrossRef CAS PubMed.
  5. A. R. bin Mohd Yusoff, D. Kim, H. P. Kim, F. K. Shneider, W. J. da Silva and J. Jang, Energy Environ. Sci., 2015, 8, 303–316 Search PubMed.
  6. R. A. J. Janssen and J. Nelson, Adv. Mater., 2013, 25, 1847–1858 CrossRef CAS PubMed.
  7. K. H. Hendriks, G. H. L. Heintges, V. S. Gevaerts, M. M. Wienk and R. A. J. Janssen, Angew. Chem., Int. Ed., 2013, 52, 8341–8344 CrossRef CAS PubMed.
  8. L. Dou, C.-C. Chen, K. Yoshimura, K. Ohya, W.-H. Chang, J. Gao, Y. Liu, E. Richard and Y. Yang, Macromolecules, 2013, 46, 3384–3390 CrossRef CAS.
  9. R. S. Ashraf, I. Meager, M. Nikolka, M. Kirkus, M. Planells, B. C. Schroeder, S. Holliday, M. Hurhangee, C. B. Nielsen, H. Sirringhaus and I. McCulloch, J. Am. Chem. Soc., 2015, 137, 1314–1321 CrossRef CAS PubMed.
  10. K. H. Hendriks, W. Li, M. M. Wienk and R. A. J. Janssen, J. Am. Chem. Soc., 2014, 136, 12130–12136 CrossRef CAS PubMed.
  11. D. Veldman, S. C. J. Meskers and R. A. J. Janssen, Adv. Funct. Mater., 2009, 19, 1939–1948 CrossRef CAS.
  12. G. W. P. Van Pruissen, F. Gholamrezaie, M. M. Wienk and R. A. J. Janssen, J. Mater. Chem., 2012, 22, 20387–20393 RSC.
  13. M. Karakawa and Y. Aso, Macromol. Chem. Phys., 2013, 214, 2388–2397 CrossRef CAS.
  14. A. P. Zoombelt, M. Fonrodona, M. M. Wienk, A. B. Sieval, J. C. Hummelen and R. A. J. Janssen, Org. Lett., 2009, 11, 903–906 CrossRef CAS PubMed.
  15. T. T. Steckler, P. Henriksson, S. Mollinger, A. Lundin, A. Salleo and M. R. Andersson, J. Am. Chem. Soc., 2013, 136, 1190–1193 CrossRef PubMed.
  16. J. Fan, J. D. Yuen, M. J. Seifter, J.-H. Seo, A. R. Mohebbi, D. Zakhidov, A. Heeger and F. Wudl, Adv. Mater., 2012, 24, 2186–2190 CrossRef CAS PubMed.
  17. E. Wang, L. Hou, Z. Wang, S. Hellström, W. Mammo, F. Zhang, O. Inganäs and M. R. Andersson, Org. Lett., 2010, 12, 4470–4473 CrossRef CAS PubMed.
  18. T. L. D. Tam, T. Salim, H. Li, F. Zhou, S. G. Mhaisalkar, H. Su, Y. M. Lam and A. C. Grimsdale, J. Mater. Chem., 2012, 22, 18528–18534 RSC.
  19. E. J. Zhou, J. Z. Cong, K. Hashimoto and K. Tajima, Energy Environ. Sci., 2012, 5, 9756–9759 CAS.
  20. S. Cho, J. Lee, M. H. Tong, J. H. Seo and C. Yang, Adv. Funct. Mater., 2011, 21, 1910–1916 CrossRef CAS.
  21. K. H. Hendriks, W. Li, G. H. L. Heintges, G. W. P. van Pruissen, M. M. Wienk and R. A. J. Janssen, J. Am. Chem. Soc., 2014, 136, 11128–11133 CrossRef CAS PubMed.
  22. D. Dang, W. Chen, R. Yang, W. Zhu, W. Mammo and E. Wang, Chem. Commun., 2013, 49, 9335–9337 RSC.
  23. W. Sun, Z. Ma, D. Dang, W. Zhu, M. R. Andersson, F. Zhang and E. Wang, J. Mater. Chem. A, 2013, 1, 11141–11144 CAS.
  24. A. P. Zoombelt, S. G. J. Mathijssen, M. G. R. Turbiez, M. M. Wienk and R. A. J. Janssen, J. Mater. Chem., 2010, 20, 2240–2246 RSC.
  25. F. L. Zhang, J. Bijleveld, E. Perzon, K. Tvingstedt, S. Barrau, O. Inganäs and M. R. Andersson, J. Mater. Chem., 2008, 18, 5468–5474 RSC.
  26. W. Li, K. H. Hendriks, A. Furlan, W. S. C. Roelofs, M. M. Wienk and R. A. J. Janssen, J. Am. Chem. Soc., 2013, 135, 18942–18948 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4cc10357b

This journal is © The Royal Society of Chemistry 2015